Polypyrrole-Decorated Ag-TiO2 Nanofibers Exhibiting Enhanced

Jun 17, 2013 - Bingxi Yan , Yongchen Wang , Xinyi Jiang , Kaifeng Liu , and Liang Guo ..... Tianyong Wang , Qing Xu , Leqiang Shao , Deli Jiang , Min ...
0 downloads 0 Views 8MB Size
Subscriber access provided by Linköpings universitetsbibliotek

Article

Polypyrrole-decorated Ag-TiO2 nanofibers exhibiting enhanced photocatalytic activity under visible light illumination Yucheng Yang, Junwei Wen, Jianhong Wei, Rui Xiong, Jing Shi, and Chun-Xu Pan ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/am401167y • Publication Date (Web): 17 Jun 2013 Downloaded from http://pubs.acs.org on June 26, 2013

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Polypyrrole-decorated Ag-TiO2 nanofibers exhibiting enhanced photocatalytic activity under visible light illumination Yucheng Yang, Junwei Wen, Jianhong Wei*, Rui Xiong, Jing Shi, Chunxu Pan*

Key Laboratory of Artificial Micro- and Nano-structures of Ministry of Education and School of Physics and Technology, Wuhan University, Wuhan 430072, China

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT:

Page 2 of 27

In the work, a novel photocatalyst, polypyrrole (PPy)- decorated Ag-TiO2

nanofibers (PPy-Ag-TiO2) with core-shell structure was successfully synthesized by electrospinning technique, followed by a surfactant directed in-situ chemical polymerization method. The results show that a PPy layer was formed on the surface of Ag-TiO2 nanofiber, which is beneficial for protecting Ag nanoparticles from being oxidized. Meanwhile, the PPyAg-TiO2 system exhibits a remarkable light absorption in the visible region and high photocurrent. Among them, the 1%- PPy-Ag-TiO2 sample shows the highest photoactivity which is far exceeded to those of the single- and two-component systems. This result may be due to the synergistic effect of Ag, PPy and TiO2 nanostructures in the ternary system.

Keywords: TiO2-based nanocomposites; silver; conducting polymer; nanofibers; core-shell structures; photocatalysis.

ACS Paragon Plus Environment

2

Page 3 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

1 Introduction Since the discovery of carbon nanotubes1, one-dimensional nanomaterials have been considered as attractive candidates for device applications because of their dimensional confinement and structurally well-defined physical and chemical properties. TiO2 nanofibers or nanotubes had attracted much attention for their potential application in the field of pollutant degradation and solar energy conversion, including dye-sensitized solar cells, photocatalyst, and hydrogen energy business2-9. However, its wide band gap (3.2 eV) and the low quantum yield largely limited the overall photocatalytic efficiency. To extend the photo response of TiO2 to the visible region, much effort was made by doping with metal or non-metal ions, sensitization with organic dyes, conducting polymer or functional carbonaceous materials10-15, etc. Especially, many reports have shown that TiO2 surfaces modified with conducting polymers such as polyaniline, polypyrrole, polythiophene and their derivatives can greatly enhanced photocatalytic activity for the degradation of organic compounds under visible light irradiation16-22. Our group has also reported that PANI sensitized TiO2 photocatalyst exhibited a good photocatalytic activity23. Coupling TiO2 with conducting polymer can efficiently promote photoinduced electron - hole pair separation in heterojunction photo-reciprocal transfer of electrons or holes, thus improving its photocatalytic activity24-25. Meanwhile, how to solve the problem of considerable recombination of the photogenerated electron–hole pairs is another challenge in enhancing the photocatalytic activity. Noble metals have been proven to be good materials for inhibiting electron–hole pair recombination by (1) increasing charge separation within the semiconductor particle, (2) discharging photogenerated electrons across the interface, and (3) providing a redox pathway with low overpotential26. Among these noble metals, Ag is a popular choice because of its relatively cheap, antibacterial

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 27

property, high work function, and ability to generate surface plasmons at the desired wavelength27. However, small Ag nanoparticles are very reactive and are easily oxidized and lost. It is therefore beneficial to have a core/shell structure to protect the Ag nanoparticles from being oxidized. The newly emerging nanotechnology offers another effective solution to improve the photocatalytic activity of TiO228-30. Choi et al31 pointed out that TiO2 nanofibers have a far more efficient charge separation/transfer process and/or recombination inhibition mechanism than TiO2 nanoparticles. The main reason is that TiO2 nanoparticles are three-dimensionally interconnected in TiO2 nanofibers. A very fast vectorial transport of photogenerated charge carriers (electrons and holes) between the particles may occur in the grain boundaries, resulting in a high photocurrent and photodegradation efficiency. What is more, TiO2 nanofibers are very hard and strong and thus can maintain its high activity in a variety of conditions even for long light irradiation periods. Although TiO2 nanoparticles have mesopore as well, they ease to aggregate in their loose and random states, which lower their photocatalytical efficiency. Based on the aforementioned considerations, constructing an artificial multi-component photocatalytical system composed of TiO2, conducting polymer and Ag with nanofiber structure is a good strategy for narrowing the band gap and increasing the quantum yield of TiO2, thereby resulting in enhanced photocatalytic activity. In this work, we describe an efficient way to synthesis of PPy- decorated Ag-TiO2 nanofibers (PPy-Ag-TiO2) with core/shell structure. Although PPy-Ag-TiO2 thin films have been reported32, studies on PPy-Ag-TiO2 nanofibers are limited. Compared with the corresponding single- and two-component samples, the three-component PPy-Ag-TiO2 system

ACS Paragon Plus Environment

4

Page 5 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

exhibits enhanced photocatalytic activity in the decomposition of gaseous acetone under visible light irradiation, which may be due to their high visible light-gathering ability, fast charge transfer rate, and low electron-hole recombination based on the photo-synergistic effect of TiO2, Ag, and PPy in the PPy-Ag-TiO2 system. Moreover, the PPy-Ag-TiO2 nanofibers can easily be recycled without decreasing the photocatalytic activity because of the large length-to-diameter ratio of the one dimensional nanostructure.

2 Experimental 2.1 Electrospinning of Ag-TiO2 heterostructure nanofibers Ag-TiO2 heterostructure nanofibers were prepared according to the reference33 with a little modification. In a typical procedure, 1 g of tetrabutyl titanate (TBT) was dissolved into a mixture, which is composed of 40 mL ethanol and 10 mL acetic acid. After the mixture was stirring for 1h, 4 g of polyvinylpyrrolidone (PVP) was slowly added into the solution. The aforementioned solution was then mixed with a silver nitrate solution (containing 0.1M AgNO3 and 0.1M sodium bis (2-ethylhexyl) sulfosuccinate (AOT) / cyclohexane, the mole ratio of AgNO3 to Ti (OC4H9)4 was 5% ). The solutions were homogeneously stirred for 10 min, and the reducing agent (containing 0.2 M NaBH4 and 0.1 M AOT/cyclohexane) was then added dropwise to the above solution still with continuous stirring. In a typical procedure for electrospinning, the precursor solution was ejected from the plastic syringe (stainless steel needle with 0.4 mm inner diameter). The metallic needle was connected to a high-voltage power supply, and a piece of aluminum foil was placed 10.0 cm below the tip of the needle to collect the nanofibers. The voltage was varied between 10.0 and 12.0 kV, the feeding rate was 1 mL / h. The as-spun nanofibers were then calcined at 500 oC for 3 h in air before cooling to room temperature. Pure TiO2 nanofibers were

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 27

similarly prepared but without the addition of AgNO3 solution and the reducing agent. The electrospinning setup consists of three major components: a high-voltage power supply (5 kV to 30 kV), a spinneret (a metallic needle), and a collector ( a ground conductor). The photograph and schematic of the experimental setup for electrospinning can be found in Figure S1 and S2 of supporting information. 2.2 Preparation of PPy-decorated Ag-TiO2 nanofiber Pyrrole (Aldrich) was distilled in a vacuum prior to use, and was immediately used, or refrigerated in air in the dark. In a typical experiment, the obtained Ag-TiO2 nanofibers were redispersed in a solution containing pyrrole (1.5 mL, 5 mM) and surfactant sodium dodecyl sulfonate ( SDS: 0.2 mL, 40 mM) fitted with ultrasonic vibration. SDS can prevent Ag-TiO2 nanofiber aggregation when pyrrole is oxidized and undergoes seeded polymerization. After 15 min, FeCl3·6H2O (1.5 mL, 5 mM) aqueous solution was added to the mixture, and ultrasonic vibration was continued for another 0.5 h. The reaction mixture was then incubated at room temperature for 12 h to ensure complete reaction. Finally, the precipitate was centrifuged, washed with deionized water and ethanol several times, and then dried in vacuum at 80 ºC until a constant mass was reached. The PPy doping concentration (X) was changed from 0.5 wt% to 2.0 wt%, and the corresponding photocatalysts were named as X%-PPy-Ag-TiO2. For comparison, PPy-TiO2 nanofibers were prepared under similar conditions. In the study, the content of PPy in PPy-TiO2 and PPy-Ag-TiO2 was 1% unless otherwise stated.

2.3 Characterization The phases of the samples were characterized by X-ray diffraction (XRD), employing a scanning rate of 0.05o per second in a 2θ ranging from 10 to 80o, using a Bruker D8 Advance X-

ACS Paragon Plus Environment

6

Page 7 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

ray diffractometer (Cu Kα, λ = 1.54178Å). The FT-IR spectra of samples were recorded with a Shimadzu IR Prestige-21. The morphologies of the samples were studied by a SHIMADZU SSX-550 field emission scanning electron microscopy (SEM), and a JEOL JEM-2010 transmission electron microscopy (TEM). Selected area electron diffraction (SAED) pattern was used to determine the composition of the samples. The UV–vis DRS was performed at room temperature on VARIAN Cary-5000 from 200nm to 800 nm, using BaSO4 as the reflectance standard. The photoluminescence (PL) spectra were recorded by F-4600 fluorescence spectrophotometer (Hitachi, Japan) under ambient condition. The excitation wavelength was 315 nm, the scanning speed was 1200 nm/min, and the slot widths of excitation slit and emission slit were both 5.0 nm. 2.4 Photoelectrochemical measurements The photocurrent developed by irradiating the photo-anode (TiO2) with either UV or visible light was recorded with an electrochemical workstation (CHI660A, CH Instruments Co.). The photoelectrochemical cell was a three-electrode system: a TiO2 film located in the middle of the cell as working electrode, a saturated calomel electrode as reference, and a platinum wire parallel to the working electrode as a counter electrode. The photo-anode was exposed to visible light to measure both open-circuit, photovoltage and closed-circuit photocurrent. The light source was a 160 W high-pressure mercury lamp with a UV cutoff filter (>420nm). All measurements were conducted at room temperature and N2 atmosphere to obtain highly reproducible data. The electrolyte was 0.5 mol/L Na2SO4 aqueous solution. The working electrode was activated in the electrolyte for 2 h before measurement. The working electrode potentials were located at 0 V to simulate the same working condition as that of the photocatalysis reaction system.

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 27

2. 5 Evaluation of photocatalytic activity Gaseous acetone was used as target substrates for the photocatalytic activity test. The photodegradation experiments were carried out in a 8L reactor under ambient condition where a 125W high-pressure mercury lamp with a 400 nm cut-off filter was used as a visible light source. 0.5 g photocatalyst powder was used for each experiment. Prior to photoreaction, the experiment was carried out in darkness for 2h to establish adsorption–desorption equilibrium. The concentrations of acetone, carbon dioxide, and water vapor were determined using a gas chromatograph. The photocatalytic activity of the samples was quantitatively evaluated according to the equation: ln (C0/C) = kt. Here, C0 and C represent the initial equilibrium concentration and reaction concentration of acetone, respectively; k represents the apparent reaction rate constant, and t represents reactive time.

3 Results and discussion Figure 1 shows the XRD patterns of the different samples. As shown in curve a, the weak reflection centered at a 2θ value of 24 ° was characteristic of the doped PPy34. The structure of the as-prepared TiO2 nanofibers was compared with the Joint Committee for Powder Diffraction Set (JCPDS) data No.21-1272 and was found to be in anatase form (curve b). A comparison of curves c and d illustrated that the Ag-TiO2 and PPy-Ag-TiO2 had similar patterns; both anatase and the metallic Ag phases (JCPDS cards No.89-3722) were detected in their patterns. However, the relative intensity of the TiO2 and Ag phases decreased in PPy-Ag-TiO2 indicating their encapsulation by PPy. FT-IR measurement results further confirmed the existence of PPy in the composite. ( Figure S3 and corresponding analysis )

ACS Paragon Plus Environment

8

Page 9 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 1. XRD patterns of (a) PPy; (b) TiO2; (c) Ag-TiO2; (d) PPy-Ag-TiO2.

The controlled morphology of the as-prepared TiO2 and Ag-TiO2 nanofibers are shown in Figure 2. Figure 2a presents the typical SEM images of the TiO2 nanofibers. The diameters of the TiO2 nanofibers are about 50 nm to 200 nm, and the length reached a few millimeters. As shown in a TEM image of the TiO2 nanofibers (Figure 2b), a single nanofiber was composed of a lot of nanoparticles, which agglomerated to form TiO2 nanofibers during the electrospinning process28. The TiO2 nanocrystals were found to be in anatase form according to the SAED pattern (inset of Figure 2b); thus it should be pure TiO2. The black dots in the TiO2 nanofibers maybe due to the fact that the orientation of the particles was close to Bragg's condition. Consequently, their diffraction was stronger than their transmission, which in turn resulted in a darker diffraction contrast image. The Ag-TiO2 nanofibers (Figure 2c) had a similar size and length with TiO2

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 27

nanofibers but only with some small dots sticking onto the surface (Figure 2d). This similarity was likely due to the use of the same precursor and synthesis route only in the presence of AgNO3 solution. According to a typical TEM image of Ag-TiO2 nanofibers (Figure S4), the average particle size of Ag was determined to be 18.02 nm (see Figure S5).

Figure 2. (a) SEM image of TiO2 nanofibers, (b) TEM image of TiO2 nanofibers (inset: corresponding SAED pattern), (c) TEM image of Ag-TiO2 nanofibers (inset: corresponding SAED pattern), (d) a high resolution TEM image of Ag-TiO2 nanofiber.

ACS Paragon Plus Environment

10

Page 11 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 3. TEM images of (a) PPy-TiO2 nanofibers, (b) a magnified PPy-TiO2 nanofiber image, (c) PPy-Ag-TiO2 nanofibers, (d) a magnified image of PPy-Ag-TiO2 nanofiber. The controlled morphology of the as-prepared PPy-TiO2 and PPy-Ag-TiO2 samples are shown in Figure 3. The TEM images revealed the PPy-TiO2 composites had a uniform smooth surface and core-shell morphology (Figures 3a and 3b). The outer layer was apparently PPy, and the inner layer was TiO2 fibers; the thickness of PPy was about 10 nm. Figures 3c and 3d show the TEM images of the PPy-Ag-TiO2 nanofibers. Compared with PPy-TiO2 nanofibers, the PPyAg-TiO2 nanofibers were rather uneven. In the two systems, SDS played important role in the formation of PPy-TiO2 nanofibers. During polymerization, SDS molecules were first absorbed onto the surface of the TiO2 (or Ag-TiO2) nanofibers. A columnar micro region containing a hard

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 27

core (TiO2 and/or Ag) and a soft interface (SDS) was then formed. After adding the pyrrole monomers and the FeCl3 oxidant, polymerization occurred between the surfactant layer and the Ag or TiO2 surface. PPy was gradually deposited onto the TiO2 (or Ag-TiO2) nanofibers surface, to form PPy -TiO2 or PPy-Ag-TiO2 nanofibers35-36. The thin layer of PPy on the surface substantially protected Ag from being oxidized.

Figure 4 UV–vis absorption spectra of different samples The UV–vis diffuse reflectance spectra (DRS) of the different samples are illustrated in Figure 4. The pure TiO2 sample showed the typical absorption of anatase with an intense transition in the UV region of the spectrum which was due to electron promotion from the valence band to the conduction band. The Ag-TiO2 sample, clearly showed a characteristic

ACS Paragon Plus Environment

12

Page 13 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

absorption of TiO2 in the UV region and a new absorption shoulder at 400 nm to 600 nm that can be attributed to the Ag surface plasmon resonance with the TiO2 interband transition at λ < 380 nm. Meanwhile, a decrease in the band gap values can be observed for Ag-TiO2 photocatalysts, as reported by other authors37-38. This decrease was due to the fact that the metallic clusters introduced localized energy levels in the TiO2 band gap. The electrons can be excited with a lower energy from the valence band (VB) to these levels rather than to the conduction band (CB) of the semiconductor. For the PPy-TiO2 and PPy-Ag-TiO2 samples, the introduction of PPy significantly affected the light absorption of TiO2 and Ag-TiO2, and the absorption intensity increased with increased PPy doping content, probably because PPy had strong absorption ability within the UV and visible wavelength range. Figure 4 shows that the presence of PPy in the materials is expressed by the continuous band from 400 nm to 800 nm with increased absorption towards the wavelength characteristic of black solids. The adsorption strength of PPy-Ag-TiO2 was markedly higher than that of PPy - TiO2 because of the existence of Ag. The PL emission mainly resulted from the recombination of excited electrons and holes, and a lower PL intensity indicated a higher separation efficiency39. The PL measurement results for different samples are presented in Figure 5. Figure 5a shows that a strong peak located at ca. 380 nm can be attributed to the recombination of free electrons from the CB bottom to the recombination center at the ground state because its energy was nearly equal to the band gap (3.20 eV) of TiO2. The other three peaks observed within the wavelength range of 390 nm to 450 nm were attributed to excitonic PL, which mainly resulted from surface oxygen vacancies and defects40. Compared with pure TiO2 sample, the PL intensity of the other three samples significant decreased, indicating that they have lower recombination rate of the photoelectrons carriers than that of pure TiO2 samples under UV light irradiation. Among these samples, PPy-

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 27

Ag-TiO2 had the lowest PL intensity, indicating that it had the lowest recombination rate of photoelectrons carriers. This result was due to the fact that the electrons were excited from the valence band to the conduction band and then transferred to the Fermi level of Ag, thereby preventing direct recombination of electrons and holes.

Figure 5. PL emission spectra (excited at 325 nm) of (a) TiO2, (b) Ag-TiO2; (c) PPy-TiO2; (d) PPy-Ag-TiO2.

ACS Paragon Plus Environment

14

Page 15 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 6 Photocurrent transient responses at a constant potential of 0.5V for (a) TiO2, (b) AgTiO2; (c) PPy-TiO2; (d) PPy-Ag-TiO2. Another methodology was used to detect the transient photocurrent responses to provide further evidence of electron–hole transfer mechanism. Figure 6 shows the typical photocurrent (I) - time (t) response curves for different samples with several on-off cycles of intermittent visible light irradiation. The initial anodic photocurrent spike caused by the separation of the electronhole pairs by movement of holes towards the semiconductor surface where they were trapped or reduced by the species in the electrolyte, whereas the electrons were transported to the back contact. After achieving the anodic photocurrent spike, the photocurrent continuously decreased with time until a steady-state photocurrent was reached. The photocurrent decay indicated that

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 27

charge recombination processes were occurring41-44. The photocurrent of the undoped TiO2, PPyTiO2, Ag-TiO2, and PPy-Ag-TiO2 electrodes were 0.18, 0.43, 0.54 and 0.83mA/cm2, respectively. The photocurrent of the Ag-TiO2, and PPy-TiO2 electrode were about 2.38, and 3 times as high as those of the TiO2 electrode, and PPy-Ag-TiO2 increased the photocurrent further to 4.6 times that of the TiO2 electrode. The photocurrent followed the order: PPy-AgTiO2 > PPy-TiO2 > Ag-TiO2 > pure TiO2.

The obvious enhancement of PPy-Ag-TiO2 in

photocurrent indicated smaller recombination and more efficient separation of photogenerated electron–hole pairs at its interface. The result agreed well with the PL measurement. The photocatalytic activity of different samples was evaluated by measuring the timedependent degradation of gaseous acetone under visible-light irradiation. The results are shown in Figure 7. It can be seen that the pure TiO2 nanofiber sample showed poor photocatalytic activities in visible light range, the degradation rate constant k was about 0.009/min, which was due to the large band gap energy of TiO2 (3.0 eV for rutile and 3.2 for anatase). The activity of Ag-TiO2 sample was much higher than that of TiO2, and its rate constant k reaches 0.023 /min. This result maybe due to the fact that Ag nanoparticles on the TiO2 surface can act as a sink for electrons, which contributed to the interfacial charge transfer between the metal and semiconductor and to the separation of photogenerated electron-hole pairs, thereby enhancing the photocatalytic activity. After introducing 1.0 wt% of PPy, the activities of PPy-TiO2 and PPyAg-TiO2 remarkably increased. The rate constant k of PPy-TiO2 is 0.048/min, which was 5.33 times that of TiO2 and 2.07 times that of Ag-TiO2. PPy-Ag-TiO2 exhibited much higher photocatalytic activity than the above samples, and its k value was 0.087/min, which was 9.66, 3.78, and 1.81 times that of TiO2, Ag-TiO2, and PPy-TiO2, respectively.

ACS Paragon Plus Environment

16

Page 17 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 7 (a) The visible-induced photocatalytical activity of different samples, (b) Relationship between photodegradation rate constant and loaded PPy content by PPy-TiO2 or PPy-Ag-TiO2.

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 27

The PPy content significantly influenced the photodegradation of gaseous acetone (Figure 7b). The photocatalytic activity of PPy-TiO2 and PPy-Ag-TiO2 initially increased and then decreased with increased PPy content from 0.5 wt% to 2.0 wt%. The optimum doping content for PPy was 1.0 wt%, which maybe due to the balance between the increase in highest occupied molecular orbital (HOMO)-electrons (PPy) potential and the decrease in light adsorption. High PPy load prevented TiO2 from absorbing visible light, consequently resulted in a rapid decrease of irradiation passing through the reaction system. The higher photocatalytic activity of 1%PPy-Ag-TiO2 may be due to the synergistic effect of Ag, PPy and TiO2 nanostructures in the ternary system. Kinetics parameters of different samples for degradation of gaseous acetone under visible light irradiation can be found in Table 1. Table 1 Kinetics parameters of different samples for degradation of gaseous acetone under visible light irradiation

Photocatalyst

Visible light K (/min)

R

TiO2

0.009

0.9958

5% Ag/TiO2

0.023

0.9903

0.5%-PPy-TiO2

0.026

0.9943

1.0 %-PPy/TiO2

0.048

0.9970

1.5%-PPy-TiO2

0.035

0.9920

2.0%-PPy-TiO2

0.030

0.9926

0.5%-PPy- Ag/TiO2

0.052

0.9984

1.0%-PPy- Ag/TiO2

0.087

0.9995

1.5%-PPy- Ag/TiO2

0.063

0.9964

2.0%-PPy- Ag/TiO2

0.055

0.9986

ACS Paragon Plus Environment

18

Page 19 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Scheme 1 Postulate mechanism of the visible light-induced photo-degradation of acetone with PPy-Ag-TiO2 nanocompsosites A schematic of the charge transfer processes of PPy -Ag-TiO2 is illustrated in Scheme 1. For PPy-TiO2, when the PPy shell harvested visible light, an absorbed photon promoted an electron from the ground state of the polymer located in the semiconductor energy gap into an excited state that was in resonance with the CB. The polymer π-orbital became the HOMO in the combined system. Given that the lowest unoccupied molecular orbital (LUMO) levels of the polymer were energetically higher than the conduction band edge of TiO245-46, the electrontransfer paths in Scheme 1 were possible. As a result, rapid charge separation and slow charge recombination occurred, resulting in increased photocatalytic activity. For PPy-Ag-TiO2, when the PPy-Ag-TiO2 composites were illuminated under visible light, their electrons can be excited from the HOMO to the LUMO of PPy, whereas holes were left in

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 27

the HOMO of PPy. The excited-state electrons can be readily injected into the CB of TiO2, and then further injected into the Fermi level of Ag, or maybe directly injected into the Fermi level of Ag. The metallic Ag nanoparticles functioned as an electron sink to accept the photogenerated electrons from the excited semiconductor, thereby facilitating dioxygen reduction47. As a result, PPy-Ag-TiO2 had a quicker charge separation and slower charge recombination process than PPy-TiO2 and thus had higher photocatalytic activity. When the addition value of PPy was higher than 1 wt%, the presence of a large amount of PPy can cover the surface of AgTiO2 and form a relatively thick layer that hindered the injection of excited electrons from the outer PPy layer to the inner TiO2 layer. Consequently, ·OH radicals decreased and the photodecomposition of the target contamination was affected. Acetone conversion obtained after five successive reaction cycles on 1%-PPy-Ag-TiO2 samples is shown in Figure 8. Catalytic recycling studies were carried out by recovering the used catalysts samples after 160 min of reaction and reusing them with fresh reagents in the subsequent reaction cycles, which was repeated five times. The recovered catalysts were washed with acetone/ethanol and dried in air before being reused in the next catalytic test. The results showed that the activity of PPy-Ag-TiO2 catalysts decreased by about 10% upon five recycling tests. The slight decrease after each cycle was attributed to the absorption of contamination and the decrease in active spots. Although slight decrements in photocatalytic activity were observed, the 1%-PPy-Ag-TiO2 particles still maintained a high level of activity in successive reusing experiments, indicating that the PPy-Ag-TiO2 prepared in this study was stable and effective for the removal of organic pollutants.

ACS Paragon Plus Environment

20

Page 21 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 8. Recycle of PPy-Ag-TiO2 under visible light irradiation

4 Conclusions In summary, PPy-Ag-TiO2 core/shell nanofibers were successfully prepared through an efficient route. The novel photocatalysts showed obvious visible light photocatalytic activity in the decomposition gaseous acetone, the 1%- PPy-Ag-TiO2 sample provided the optimum photocatalytic activity compared with the pure TiO2 nanofibers, Ag-TiO2 nanofibers, and PPyTiO2 core-shell nanofibers under visible-light irradiation. The high photoactivity of the PPy-AgTiO2 can be attributed to the synergistic effect originating from the excited-state electrons in PPy can be readily injected into the TiO2 CB and be further injected into the Fermi level of Ag. As a result, rapid charge separation and slow charge recombination occurred, resulting in increased photocatalytic activity. The recycling test revealed that the PPy-Ag-TiO2 prepared in this study was stable and effective for the removal of organic pollutants. Therefore, the method described

ACS Paragon Plus Environment

21

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 27

in this paper provided a simple and effective strategy for the rational design of delicate composite photocatalysts for applications beyond photocatalysis. AUTHOR INFORMATION Correponding authors

*:E-mail: [email protected](J. H. Wei); [email protected] (C. X. Pan).

Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS

We are grateful for the financial support from the National Natural Science Foundation of China (No. 51272185) and the National Program on Key Basic Research Project (973 Grant No. 2009CB939705 & 2012CB821404). ASSOCIATED CONTENT Supporting information Available The photograph and schematic of the basic experimental setup for electrospinning, FT-IR spectra of different samples, the diagram of Ag particle size distribution, photocurrent transient responses, and UV-Vis spectra of PPy can be found in supporting information. This information is available free of charge via the internet at http://pubs.acs.org.

ACS Paragon Plus Environment

22

Page 23 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

REFERENCES: (1) Iijima, S. Nature 1991, 354, 56-58. (2) Tong, H.; Ouyang, S. X.; Bi, Y. P.; Umezawa, N.; Oshikiri, M. ; Ye, J. H. Adv. Mater. 2012, 24, 229-251. (3) Zhou, W. J.; Du, G. J.; Hu, P . G.; Wang, D. Z.; Liu, H.; Wang, J. Y. ; Boughton, R. I. ; Liu, D. ; Jiang, H. D. J. Mater. Chem. 2011, 21,7937-7945. (4) Choi, S. K.; Kim, S.; Lim, S. K.; Park, H. J. Phys. Chem. C 2010, 114,16475-16480. (5) Shah, M.S.A. S.; Park, A. R.; Zhang, K.; Park, J. H.; Yoo, P. J. ACS Appl. Mater. Interfaces 2012, 4, 3893-3901. (6) Yang, L; Xiao, Y; Liu, S; Li, Y; Cai, Q; Luo, S; Zeng, G. Appl. Catal. B 2010, 94, 142-149. (7) Tang, Y; Luo, S; Teng, Y; Liu, C; Xu, X; Zhang, X; Chen, L. J. Hazard. Mater. 2012, 241242, 323-330. (8) Yang, L; Chen, B; Luo, S; Li, J; Liu, R; Cai, Q. Environ. Sci. Tech. 2010, 44, 7884-7889. (9) Yang, L; Luo, S; Li, Y; Xiao, Y; Kang, Q; Cai, Q. Environ. Sci. Tech. 2010, 44, 7641-7646. (10) Livraghi, S.; Paganini, M. C.; Giamello, E.; Selloni, A.; Valentin, C. D.; Pacchioni, G. J. Am. Chem. Soc. 2006, 128, 15666-15671. (11) Mohamed, A. E. R.; Rohani, S. Energy Environ. Sci. 2011, 4,1065-1086. (12) Lee, S. S.; Fan, C. Y.; Wu, T. P.; Anderson, S. L. J. Am. Chem. Soc. 2004, 126, 5682-5683. (13) Zhang, L. W.; Fu, H. B.; Zhu, Y. F. Adv. Funct. Mater. 2008, 18, 2180-2189. (14) Cheng, H.; Zhao, X. J.; Sui, X. T.; Xiong, Y. L.; Zhao, J. J. Nanopart. Res. 2011, 13, 555562.

ACS Paragon Plus Environment

23

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 27

(15) Chuang, H. Y.; Chen, D. H. Nanotechnology, 2009, 20,105704/1-105704/10. (16) Kandiel, T. A.; Dillert, R.; Bahnemann, D. W. Photochem. Photobiol. Sci. 2009, 8, 683-690. (17) Chen, A. H.; Xie, H. X. ; Wang, H.Q.; Li, H. Y. ; Li, X. Y. Synth. Met. 2006, 156, 346-350. (18) Chowdhury, D. ; Paul, A.; Chattopadhyay, A. Langmuir 2005, 21, 4123- 4128. (19) Xu, S. B.; Zhu, Y. F.; Jiang, L. ; Dan, Y. Water, Air, Soil Pollut. 2010, 213,151-159. (20) Li, X. Y.; Wang, D. S.; Cheng, G. X.; Luo, Q. Z.; An, J.; Wang, Y. H. Appl. Catal. B 2008, 81, 267-273. (21) Li, S. Y. ; Chen, M. K.; He, L. J.; Xu, F.; Zhao, G. H. J. Mater. Res. 2009, 24, 2547-2554. (22) Huang, K.; Wan, M. X. ; Long,Y. Z.; Chen, Z. J. ; Wei, Y. Synth. Met. 2005, 155, 495-500. (23) Wei, J. H.; Zhang, Q.; Liu, Y.; Xiong, R.; Pan, C. X.; Shi, J. J. Nanopart. Res. 2011,13, 3157-3165. (24) Frank A. J.; Honda, K. J. Photochem. 1985, 29, 195-204. (25) Wang, J.; Ni, X. Y. Solid State Commun. 2008, 146, 239-244. (26) Kamat, P. V. J. Phys. Chem. Lett. 2012, 3, 663-672. (27) Kamat, P. V. J. Phys. Chem. B 2002, 106, 7729-7744. (28) Zhang, P.; Shao, C. L.; Zhang, Z. Y.; Zhang, M. Y.; Mu, J. B.; Guo, Z. C.; Sun,Y. Y.; Liu, Y. C. J. Mater. Chem. 2011, 21, 17746-17753. (29) Chuangchote, S.; Jitputti, J.; Sagawa, T.; Yoshikawa, S. ACS Appl. Mater. Interfaces 2009, 1, 1140-1143.

ACS Paragon Plus Environment

24

Page 25 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(30) Lu, X. F.; Mao, H.; Zhang, W. J. Nanotechnology 2007, 18, 025604/1-025604/5. (31) Choi, S. K.; Kim, S.; Lim, S. K.; Park, H. J. Phys. Chem. C 2010, 114,16475-16480. (32) Su, P. G.; Chang, Y. P. Sens. Actuators, B 2008, 129, 915-920. (33) Ochandaw, F. O.; Barnett, M. R. J. Am. Ceram. Soc. 2010, 93, 2637-2643. (34) Wang, B.; Li, C.; Pang, J. F.; Qing, X. T.; Zhai, J. P.; Li, Q. Appl. Surf. Sci. 2012, 258, 9989-9996. (35) Pinter, E.; Patakfalvi, R.; Fu1lei, T.; Ging, Z.; Dekany, I.; Visy, C. J. Phys. Chem. B 2005, 109, 17474-17478. (36) Ye, S.; Fang, L.; Lu, Y. Phys. Chem. Chem. Phys. 2009, 11, 2480-2484. (37) Hu, C.; Lan, Y. Q.; Qu, J. H.; Hu, X. X.; Wang, A. M. J. Phys. Chem. B 2006, 110, 4066 4072. (38) Melian,E. P.; Diaz, O. G.; Rodriguez, J.M. D.; Colon, G.; Navio, J.A.; Macias, M.; Pena, J. P. Appl. Catal. B 2012, 127, 112-120. (39) Ishibashi, K.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Electrochem. Commun. 2000, 2, 207-210. (40) Yu, J. G.; Xiang, Q. J.; Zhou, M. H. Appl. Catal. B 2009, 90, 595-602. (41) Spadavecchia, F.; Ardizzone, S.; Cappelletti, G.; Falciola, L.; Ceotto, M.; Lotti, D. J. Appl. Electrochem. 2013, 43, 217-225. (42) Dholam, R.; Patel, N.; Santini, A.; Miotello, A. Int. J. Hydrogen Energy 2010, 35, 95819590.

ACS Paragon Plus Environment

25

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 27

(43) Hagfeldt, A.; Lindstrom, H.; Sodergren, S.; Lindquist, S. E. J. Electroanal. Chem. 1995, 381, 39- 46. (44) Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T. J. Am. Chem. Soc. 2004, 126, 5851-5858. (45) Wang, D. S.; Wang, Y. H. ; Li, X. Y.; Luo, Q. Z.; An, J.; Yue, J. X. Catal. Commun. 2008, 9, 1162-1166. (46) Murakoshi, K.; Kogure, R.; Wada, Y.; Yanagida, S. Chem. Lett. 1997, 26, 471- 472. (47) Liu, R.; Wang, P.; Wang, X. F.; Yu, H. G.; Yu, J. G. J. Phys. Chem. C 2012, 116, 1772117728.

ACS Paragon Plus Environment

26

Page 27 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Table of Contents Graphic

ACS Paragon Plus Environment

27